Published on Web 07/03/2004
Generation of Fullerenyl Cation (EtO)2P+(OH)CH2-C60+ from RC60-H and from RC60-C60R (R ) CH2P(O)(OEt)2) Yasujiro Murata, Fuyong Cheng, Toshikazu Kitagawa, and Koichi Komatsu* Institute for Chemical Research, Kyoto UniVersity, Uji, Kyoto 611-0011, Japan Received April 30, 2004; E-mail:
[email protected] It is now generally accepted that fullerenes are typically electronegative molecules, and studies on their cationic species are quite limited.1,2 Thus, there have so far been reported only two methods for the generation of monofunctionalized C60 cations as stable species;3,4 RC60+ (R ) CHCl2 or CCl2CH2Cl) was generated by ionization of the corresponding monohydroxylic C60 (fullerenol) in a strong acid such as CF3SO3H,3 and HC60+ was formed by direct protonation of C60 by an extraordinarily strong superacid having a carborane structure.4 However, fullerenol is still a rather rare material5 and the latter superacid is not commonly available. The fullerenyl cation will become a more common and widely applicable species if monofunctionalized hydrofullerenes, RC60-H, which can be synthesized by a variety of reactions,6 can be used as the precursor for the cation RC60+. Here we report the first example of the generation of a fullerenyl cation, (EtO)2P+(OH)CH2-C60+, from RC60-H7 and also from RC60-C60R (R ) CH2P(O)(OEt)2)8 as precursors. When RC60-H (2)7 was simply dissolved in H2SO4 at room temperature under air, a reddish purple solution was immediately formed. The 1H NMR spectrum (D2SO4) showed that a single species was produced with disappearance of a signal for the proton originally attached to the C60 cage in 2. As shown in Figure 1a, the UV-vis-NIR spectrum (H2SO4) showed absorptions at 488 nm (log 3.75), 795 (3.42), and ∼1122 (2.80), which are characteristic to the monofunctionalized C60 cations,3,4 the molar absorption coefficients being comparable to the reported values.3 Quenching this solution with CF3CH2OH (TFE) afforded fullerenyl ether RC60-OCH2CF3 in 85% isolated yield as a mixture of 1,2and 1,4-isomers (4 and 5, respectively) in a ratio of 4:1. These findings indicate the nearly quantitative formation of a fullerenyl cation from 2 (Scheme 1). In exactly the same way, the same cationic species was generated quantitatively from singly-bonded fullerene dimer 3. The 13C NMR spectrum (H2SO4-CF3SO3H (1:4)) of the cationic species showed 29 signals in the region between 152.05 and 136.28 ppm in addition to a signal at 53.66 ppm for the sp3 carbon on the C60 cage (Figure 1b), indicating Cs symmetry in this species. The cationic center exhibited a signal at 174.67 ppm, which is close to the reported ones (175.6 ppm for CHCl2-C60+ and 174.9 ppm for CH2ClCCl2-C60+).3 This signal appeared as a doublet, which should be due to a coupling with the phosphorus atom (JP-C ) 6.6 Hz), whereas the corresponding carbon in 2 showed no such coupling.7 Thus, it was suggested that an oxygen atom in the phosphoryl group is coordinated to the cationic center to form a five-membered cyclic structure such as A in Figure 2. In support of this, the 31P NMR spectrum (D2SO4) of this cationic species exhibited a signal at 31.08 ppm, which is downfield shifted by 7 ppm as compared with that of 2, indicative of the presence of some positive charge on the phosphorus atom. The structure of the cationic species was investigated by DFT calculations, particularly with regard to the possible coordination 8874
9
J. AM. CHEM. SOC. 2004, 126, 8874-8875
Figure 1. (a) UV-vis-NIR spectrum in H2SO4 and (b) the spectrum in H2SO4-CF3SO3H (1:4) of the cationic species.
13C
NMR
Scheme 1
of the phosphoryl group to the cationic center as well as the conformation of the phosphorylmethyl group (structures A-C in Figure 2), in H2SO4 using the Onsager model.9,10 Although A was calculated to be more stable than B and C by 12.4 and 13.0 kcal mol-1, respectively, the calculated chemical shift of C2 in A (116.88 ppm) obtained by the GIAO method11,12 was completely different from the experimental value (174.67 ppm). Since the protonation on a phosphoryl oxygen is quite possible in a strongly acidic medium,13 we next examined the structures formed by protonation on A-C. The calculations using either A or B as an initial geometry with protonation on the phosphoryl oxygen gave D as an energy10.1021/ja047483h CCC: $27.50 © 2004 American Chemical Society
COMMUNICATIONS
one-electron oxidation with the aminium salt. The formation of dimer 3 confirmed the proposed mechanism. In summary, a novel fullerenyl cation 1+ was generated in sulfuric and sulfonic acids as well as in CH2Cl2 from monofunctionalized hydrofullerene 2 and from singly bonded fullerene dimer 3 through one-electron oxidation as a key step. Various nucleophiles, including nonactivated benzene,18 readily add to 1+ selectively at 2- or 4-positions, thus showing the possibility of these reactions to be used as a new way to functionalize fullerenes.
Figure 2. Possible structures of monocationic species (A-C) and their protonated dicationic structures (D and E), including the calculated NMR chemical shifts of C2 and P obtained by the GIAO method.12
minimized structure. This structure was calculated to be more stable than nonbridged structure E by 2.8 kcal mol-1, the distance of C+‚‚‚O in D being 3.252 Å. The results of GIAO calculations on D (174.60 ppm for C2 and 26.29 ppm for P)12 were in agreement with the experimental values. Thus, we conclude that the most possible structure of the cationic species 1+ formed in H2SO4 is D with the cationic center coordinated by the protonated oxygen. For this structure, Mulliken charges on C2 and P are calculated as +0.084 and +0.658, respectively, while those of other 58 sp2 carbons of the C60 cage are in the range between +0.058 and -0.003, the sum of which amounts to +0.858.12 The rate of the formation of cation 1+ from 2 was found to depend not on the acidity but on the oxidizing ability of the acid used as a solvent; 1+ was generated immediately after dissolution in H2SO4 and FSO3H (acidity function H0, -12 and -15.1, respectively),14 whereas the generation of 1+ required more than 10 h in CF3SO3H (H0, -14.1).14,15 Therefore, the direct protonolytic cleavage of the RC60-H bond can be ruled out. When dimer 3 was used as the starting material instead of 2, cation 1+ was generated immediately in H2SO4, FSO3H, and CF3SO3H15 as well. As the reaction mechanism, we assume that first the one-electron oxidation of 2 takes place to give radical cation 2•+. Then, 2•+ would release the proton directly attached to the C60 cage to give radical 1•, which should be in equilibrium with dimer 3.8 The second oneelectron oxidation of 1• can give cation 1+, similarly to the formation of azafullerenyl cation C59N+ from its dimer.16 To confirm the above mechanism, one-electron oxidation of 2 and 3 was conducted in CH2Cl2 using a triarylaminium salt. When dimer 3 and two molar amounts of (2,4-Br2C6H3)3N•+SbF6-(HSbF6)0.5 (8),17 were mixed at room temperature for 15 min in CH2Cl2 under vacuum, a dark red solution with absorption maxima at 468 and 779 nm, with a weak and broad absorption in a NIR region, was obtained. Quenching this solution with TFE, allyltrimethylsilane, or benzene afforded the corresponding electrophilic addition product, 4, 6, or 7, in 72, 60, or 63% yield, respectively, indicating that cation 1+ was actually generated in CH2Cl2. When 2 was used as a starting material, which was less reactive than dimer 3 in oxidation in CF3SO3H, the treatment with three molar amounts of aminium salt 8 in CH2Cl2 also afforded a dark red solution. After 25 min, quenching this solution with allyltrimethylsilane gave 6 (45%) and dimer 3 (44%), while 11% of 2 remained unchanged, as judged from the 1H NMR of the crude product. The formation of 6 indicates that cation 1+ was also generated from 2 through
Acknowledgment. This work was supported by a Grant-in-Aid for COE Research for Elements Science (No. 12CE2005) from the Ministry of Education, Culture, Sports, Science and Technology, Japan. F.C. thanks Japan Society for the Promotion of Science for the Postdoctoral Fellowship for Foreign Researchers. Supporting Information Available: Experimental procedures, spectroscopic data, results of calculations (PDF), and X-ray crystallographic data for 8 (CIF). This material is available free of charge via the Internet at http://pubs.acs.org. References (1) For reviews, see: (a) Reed, C. A.; Bolskar, R. D. Chem. ReV. 2000, 100, 1075. (b) Kitagawa, T.; Takeuchi, K. Bull. Chem. Soc. Jpn. 2001, 74, 785. (2) For an example of pentaaryl derivative, Ar5C60+, see: Avent, A. G.; Birkett, P. R.; Kroto, H. W.; Taylor, R.; Walton, R. M. Chem. Commun. 1998, 2153. (3) Kitagawa, T.; Sakamoto, H.; Takeuchi, K. J. Am. Chem. Soc. 1999, 121, 4298. (4) Reed, C. A.; Kim, K.-C.; Bolskar, R. D.; Mueller, L. J. Science 2000, 289, 101. (5) There have been only a limited number of examples: (a) Yoshida, M.; Morinaga, Y.; Iyoda, M.; Kikuchi, K.; Ikemoto, I.; Achiba, Y. Tetrahedron Lett. 1993, 34, 7629. (b) Irngartinger, H.; Weber, A. Tetrahedron Lett. 1997, 38, 2075. (c) Ohno, M.; Yashiro, A.; Tsunenishi, Y.; Eguchi, S. Chem. Commun. 1999, 827. (d) ref 3. (6) (a) Hirsch, A. The Chemistry of Fullerenes; Thieme: Stuttgart, 1994. (b) Diederich, F.; Thilgen, C. Science 1996, 271, 317. (c) Taylor, R. The Chemistry of Fullerenes; World Scientific Publishing Co.: River Edge, NJ, 1996. (7) Allard, E.; Cheng, F.; Chopin, S.; Delaunay, J.; Rondeau, D.; Cousseau, J. New J. Chem. 2003, 27, 188. (8) Cheng, F.; Murata, Y.; Komatsu, K. Org. Lett. 2002, 4, 2541. (9) (a) Wong, M. W.; Frisch, M. J.; Wiberg, K. B. J. Am. Chem. Soc. 1991, 113, 4776. (b) Wong, M. W.; Frisch, M. J.; Wiberg, K. B. J. Am. Chem. Soc. 1992, 114, 523. (c) Wong, M. W.; Frisch, M. J.; Wiberg, K. B. J. Am. Chem. Soc. 1992, 114, 1645. (10) Structures were fully optimized at the B3LYP/6-31G* level in a gas phase, and then the molecular volumes were computed. By the use of the volume and the dielectric constant ( 101 for H2SO4), the structures were fully optimized at the same level of theory using the Onsager model. (11) Wolinski, K.; Hinton, J. F.; Pulay, P. J. Am. Chem. Soc. 1990, 112, 8251. (12) Based on the fully optimized structures at the B3LYP/6-31G* level of theory in a gas phase, the GIAO calculations were conducted using the B3LYP method and the basis sets 6-311G** for C, H, and O and 6-311+G(3df) for P. (13) Dillon, K. B.; Nisbet, M. P.; Waddington, T. C. J. Chem. Soc., Dalton Trans. 1981, 212. (14) Olah, G. A.; Prakash, G. K. S.; Sommer, J. Superacids; John Wiley & Sons: New York, 1985. (15) Although CF3SO3H is known as a nonoxidizing acid, (CF3SO2)2O, which is inevitably formed during the distillation of CF3SO3H, can act as an oxidizing agent: Maas, G.; Stang, P. J. J. Org. Chem. 1981, 46, 1606. (16) Kim, K.-C.; Hauke, F.; Hirsch, A.; Boyd, P. D. W.; Carter, E.; Armstrong, R. S.; Lay, P. A.; Reed, C. A. J. Am. Chem. Soc. 2003, 125, 4024. (17) Compound 8 was newly synthesized and the structure determined by X-ray crystallography. Crystal data of 8: C18H9.5Br6F9Sb1.5, M ) 1072.85, rhombohedral, space group R-3c, a ) 13.533(5) Å, c ) 51.977(5) Å, V ) 8244(4) Å3, Z ) 12, F(000) ) 5904, Dc ) 2.593 g cm-3, µ(Mo KR) 10.281 mm-1, 109 variables refined on F2 with 1618 observed reflections collected at 100 K (θmax ) 25.0°) with I g 2σ(I) yielding R1 ) 0.0675, wR2 ) 0.1690. For the same radical cation having other counteranions, see: (a) Schmidt, W.; Steckhan, E. Chem. Ber. 1980, 113, 577. (b) Bolskar, R. D.; Mathur, R. S.; Reed, C. A. J. Am. Chem. Soc. 1996, 118, 13093. (18) Anisole is reported to react with monofunctionalized C60 cations: Kitagawa, T.; Lee, Y.; Hanamura, M.; Sakamoto, H.; Konno, H.; Takeuchi, K.; Komatsu, K. Chem. Commun. 2002, 3062.
JA047483H
J. AM. CHEM. SOC.
9
VOL. 126, NO. 29, 2004 8875